1 s2.0 S0016236124007841 Main

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

Fuel 368 (2024) 131636

Contents lists available at ScienceDirect

Fuel
journal homepage: www.elsevier.com/locate/fuel

Full Length Article

Ignition characteristics of ammonia-methanol blended fuel in a rapid


compression machine
Qihang Zhang , Ridong Zhang , Yunliang Qi , Zhi Wang *
School of Vehicle and Mobility, Tsinghua University, Beijing 100084, China

A R T I C L E I N F O A B S T R A C T

Keywords: Ammonia has attracted wide attention in recent years as a carbon-free fuel. However, the low reactivity and
Ammonia combustion inertness of ammonia pose challenges to its applications in engines. Adding highly reactive fuels,
Methanol such as renewable carbon–neutral methanol, offers a potential solution. To investigate the ignition characteristics
Ignition delay time
of ammonia-methanol blended fuel, ignition delay time measurement and fast gas sampling under stoichiometric
Sampling
conditions with four ammonia blending ratios (20 % ammonia, A20; 40 % ammonia, A40; 80 % ammonia, A80;
Intermediate species
95 % ammonia, A95) were carried out on a rapid compression machine under engine-relevant conditions. The
effective thermodynamic conditions were 15–25 bar and 810–970 K. The concentrations of fuels NH3, CH3OH,
and intermediate species N2O, CO, as well as products N2, CO2 were detected using gas chromatography.
Chemical analysis was performed based on simulations using five ammonia-methanol reaction mechanisms. The
ignition delay time results showed that the addition of methanol shortened the ignition delay time, with only a
little change in ignition delay time beyond 20 % methanol content. Methanol significantly promoted the pro­
duction of OH radical, leading to the enhancement of the overall mixture reactivity. The sampling results showed
different consumption patterns of ammonia and methanol during the ignition process at different mixing ratios.
For A80, methanol was consumed from the early stage of the ignition process, while ammonia consumption was
negligible in the early stage. Conversely, for A95, ammonia consumption began in the early stage. This suggests
that methanol and its intermediate species inhibited the consumption of ammonia, however, ammonia promoted
the consumption of methanol. Chemical analysis further revealed that the inhibitory effect of methanol on
ammonia consumption was weakened with the decrease of methanol proportion.

1. Introduction methanol is a fuel that can exist in liquid form under normal atmo­
spheric conditions, which makes it easier to store and transport
In recent years, the intensification of global climate change has led to compared to methane and hydrogen [1,15]. These features facilitate the
the implementation of stricter international emission regulations and easy blending of ammonia and methanol in liquid form for convenient
the adoption of carbon reduction policies. Consequently, carbon–neutral filling and use. Moreover, methanol can be produced through carbon
fuels emerged as a promising solution [1–3], with ammonia and capture technology or biomass technology [15,16], thereby achieving
hydrogen gaining significant attention as carbon-free alternatives. carbon neutrality.
Compared to hydrogen, ammonia offers several advantages due to its The combustion of blended fuels is complicated. The investigation
mature production chain, well-developed infrastructure, lower storage conducted by Dai et al. [9] indicated that hydrogen strongly enhanced
and transportation costs, and most notably, its inherent safety in terms the ignition process of ammonia, particularly in the formation and
of being less prone to explosions [4–6]. However, the low reactivity and decomposition of hydrogen peroxide, but hydrogen did not affect the
combustion inertness of ammonia limit its use as a single fuel in engines oxidation path of ammonia. Another study of Dai et al. [8] reported a
[5,6]. To overcome this limitation, ammonia is commonly blended with similar ignition-enhancing effect of methane on ammonia. They found
highly reactive fuels such as hydrogen, methane, and dimethyl ether that ammonia was involved in methane consumption process through
(DME) [7–13]. Among highly reactive fuels, methanol is uniquely ad­ the reaction CH4 + NH2 = CH3 + NH3. Meng et al. [13] investigated the
vantageous for its large ammonia solubility [14]. Additionally, combustion and emission of ammonia/DME dual-fuel, finding that DME

* Corresponding author.
E-mail address: wangzhi@tsinghua.edu.cn (Z. Wang).

https://doi.org/10.1016/j.fuel.2024.131636
Received 7 January 2024; Received in revised form 26 March 2024; Accepted 1 April 2024
Available online 6 April 2024
0016-2361/© 2024 Elsevier Ltd. All rights reserved.
Q. Zhang et al. Fuel 368 (2024) 131636

improved the combustion progress of ammonia. The addition of DME Tsinghua University (TU-RCM), as shown in Fig. 1. An optimized crev­
increased the production of NO by yielding CH3 radicals, which also iced piston was used in the TU-RCM to eliminate temperature in­
influenced the oxidation path of ammonia. These investigations homogeneity. The premixed mixture in the combustion chamber was
demonstrate that although additives can improve the ignition perfor­ compressed within 25–30 ms. Further details on the description and
mance of ammonia, different additives can interact with ammonia in functioning of the TU-RCM can be found in Refs. [30,31].
various ways and have distinct effects on the oxidation process, leading
to complex ignition characteristics for blended fuels. Similar work can 2.2. Sampling system and gas chromatography
be found in Refs. [17–19]. Although the combustion characteristics and
engine performance of methanol blended with traditional hydrocarbon A fast gas-sampling system coupled with the RCM was used to take a
fuels have been widely studied [20–24], its blending with ammonia small portion of reacting gas and freeze the intermediate species during
remains inadequately explored. the compression ignition process of ammonia-methanol mixture. The
Previous studies have shown that the addition of methanol can sampling system comprised a sampling probe, a solenoid valve (Parker
promote the combustion of ammonia. Wang et al. [25] used the heat flux Series 9), a sampling chamber equipped with a pressure sensor (Kistler
method to measure the laminar burning velocity of ammonia-methanol 4045A), a laser trigger system (consisting of a laser sensor and a signal
mixture at 298–448 K. They found that the addition of methanol to generator RIGOL DG2052), a vacuum pump (Agilent DS 102), and
ammonia had a similar promotion effect on the laminar burning velocity vacuum tubes, as depicted in Fig. 1. Positioned to protrude 2 mm from
as hydrogen and methane. Xu et al. [18] conducted a numerical inves­ the inner wall of the combustion chamber, the sampling probe facili­
tigation on the combustion characteristics of ammonia-methanol tated sampling from the ’adiabatic core region’ of the reacting gas [32].
mixture, revealing that the addition of methanol has a strong Before each test, the sampling system was vacuumized to less than
combustion-promoting effect. Compared with pure ammonia, the peak 0.01 Torr and the solenoid valve remained closed. At predetermined
laminar burning velocity increased by 60 % when 20 % methanol was time after the EOC, the sampling solenoid valve would be opened for a
added. Li et al. [26] studied the ignition delay time of ammonia- fixed duration (4 ms in this work) under the control of the signal
methanol mixture using a shock tube at temperatures of 1250–2150 K generator. During the 4 ms solenoid valve opening, only a small amount
and pressures of 0.14 and 1.0 MPa, based on which they developed a of the mixture (less than 1 % [33]) driven by the pressure difference
simplified mechanism for chemical kinetics analysis. They concluded flowed into the sampling system, resulting in negligible changes to the
that adding 5 % methanol could shorten the ignition delay time of pressure in the combustion chamber. Fig. 2 illustrates the pressure trace
ammonia-methanol mixture by more than 60 %. The ignition delay time of a typical sampling experiment. To facilitate comparison between
of ammonia-methanol mixture was mainly affected by small radicals, sampling timings across tests with disparate ignition delay times, a
such as OH, O, HO2 and H. More studies on ammonia-methanol can be normalized sampling timing, as defined by Eq. (1), was employed. This
found in Refs. [27,28]. normalized timing is the ratio of the solenoid valve trigger timing
While the above studies provide valuable insights, their conditions τsampling to the mixture ignition delay time τign. As shown in Fig. 2,
differ from those (medium to high temperature and high pressure) τsampling and τign denote the time intervals from the EOC to the timing of
required for ammonia use in internal combustion engines. Li et al. [29] maximum pressure rise rates in the sampling chamber and combustion
used a rapid compression machine to measure the ignition delay time of chamber, respectively. It is worth noting that τ = 0 is the EOC timing,
ammonia-methanol mixtures at temperatures ranging from 845 K to and τ = 1 is the ignition timing.
1100 K and pressures of 20 and 40 bar. The experimental results
τsampling
demonstrated that the addition of methanol enhanced the reactivity of τ= (1)
τign
the mixture. Then they developed a chemical reaction mechanism for
ammonia-methanol and conducted a kinetic analysis. The results The sampled mixture obtained from each test was then analyzed by
revealed that the addition of methanol enriched the O/H radical pool gas chromatography (GC, Agilent 7890B). Ultra-high purity helium
through different reaction pathways, leading to the consumption of (>99.99 %) was used as the carrier gas for the two detectors. Fig. 3
ammonia and promoting auto-ignition. Nevertheless, their investigation presents the signal information of the species detected by the two de­
paid little attention to the interaction between ammonia, methanol and tectors. A flame ionization detector (FID) detector, operating with the
their intermediate species, which plays an important role in the ignition hydrogen/air flame at 250 ◦ C, was used to detect methanol in this study,
process of the blended fuel, as mentioned earlier. The data of species while a thermal conductivity detector (TCD), operating at 150 ◦ C, was
concentration during the ignition of ammonia-methanol mixture have employed for the detection of carbon dioxide, nitrous oxide, nitrogen,
not been reported in previous works, which will provide strong support carbon monoxide, and ammonia. Detailed description of the sampling
for the study of C-N interaction reaction. Therefore, the detailed ignition system and gas chromatograph can be found in Ref. [34].
characteristics of ammonia-methanol blended fuel needs to be further
studied. 2.3. Test conditions
Under such a background, in the current work, ignition delay time
measurement and fast gas-sampling experiments of the ammonia- Four stoichiometric ammonia-methanol/oxygen/argon mixtures
methanol blended fuel were carried out on a rapid compression ma­ were tested under various thermodynamic conditions in this work,
chine under engine-relevant conditions. Simulations using ammonia- which were given in Table 1. Based on the relative molar ratio of
methanol chemical mechanisms were also performed, based on which, ammonia and methanol in the mixture, the four kinds of mixtures were
chemical analysis was conducted. After introducing the experimental named by A20 (ammonia 20 %, methanol 80 %), A40 (ammonia 40 %,
setup and model validation, the study elucidates the ignition-enhancing methanol 60 %), A80 (ammonia 80 %, methanol 20 %), A95 (ammonia
effect of methanol on ammonia, the fuel consumption patterns of the 95 %, methanol 5 %), respectively. Argon was used as the buffer gas with
blended mixture, and the interactive reaction paths of ammonia and a dilution ratio of 8 to improve the effective temperature. In the prep­
methanol. aration of the mixtures, high purity argon (>99.99 %), oxygen (>99.99
%), methanol (>99 %), and ammonia (>99 %) were filled into a stain­
2. Experimental setup less tank in accordance with Dalton’s partial pressure law. The mixtures
were left still for more than 2 h to ensure homogeneity [35].
2.1. RCM system The initial thermodynamic conditions of the experimental mixture
were indicated by the effective pressure (peff) and the effective tem­
The experiments were conducted in a rapid compression machine at perature (Teff), taking into account the heat loss after the end of

2
Q. Zhang et al. Fuel 368 (2024) 131636

Fig. 1. Schematic of TU-RCM system.

Fig. 2. A typical pressure trace of the RCM sampling experiment.

compression (EOC). The effective pressure (peff) is defined as the average ∫ tmin
1
pressure from the first peak pressure (pmax) to the minimum pressure peff = pdt (2)
(pmin) after the EOC, as shown in Eq. (2). Fig. 2 shows that time 0 is the (tmin − tEOC ) tEOC

time of the EOC, tEOC, and the corresponding pressure is the first peak ∫ Teff ( )
pressure. Teff is defined by Eq. (3) based on isentropic correlation, in­ γ peff
dlnT = ln (3)
corporates the initial temperature (T0), initial pressure (p0), and specific T0 γ− 1 p0
heat ratio (γ).The effective pressures were adjusted to 15, 20 and 25 bar
by changing the initial pressure of the mixture charged into the com­
bustion chamber, and the effective temperatures were varied from 2.4. Simulation method
around 810 K to around 970 K by changing the length of the combustion
chamber. The zero-dimensional homogeneous reactor of CHEMKIN software
was used to calculate the ignition delay times and species information to
verify the mechanisms and do chemical analysis. Two simulation models
were applied: the constant volume (CV) model, which only considers the

3
Q. Zhang et al. Fuel 368 (2024) 131636

Fig. 3. A typical GC chromatogram of the species in the ammonia-methanol compression ignition process.

the experimental results for A80 but considerably different for A95. The
Table 1
calculated results of M4 are also very close to the experimental results
Mixture composition and test conditions.
for A80, but not as good as M3 and M5 in the other three ammonia
nNH3/(nNH3 + peff (bar) Teff (K) Equivalent ratio Dilution ratios. There is no mechanism capable of accurately predicting the
nCH3OH) ratio
ignition delay time of ammonia-methanol blends across different ratios.
φ

20 % (A20) 15/20/ 810–970 1 8 This may stem from a disconnect between applicable conditions of the
40 % (A40) 25 (Pure Ar)
mechanisms and the actual experimental conditions. Under engine-
80 % (A80)
95 % (A95) relevant conditions, the accuracy of the mechanisms in predicting the
ignition delay time of ammonia-methanol fuel in a wide range of pro­
portions needs to be improved. The conclusions drawn from the results
ideal case ignoring the heat loss during combustion, and the volume at peff = 15 bar and 20 bar are the same.
history (VH) model, which takes heat loss into account. The calculation In addition, it is noteworthy that for M3, compared with the calcu­
method for the VH model can be referred to Ref. [31]. Five mechanisms lated results using the CV model, the ignition delay times calculated by
including ammonia and methanol were used for the calculations, as using the VH model are more consistent with the experimental results
shown in Table 2. However, there are currently a limited number of only when the proportion of methanol is lower (A80, A95), while the VH
mechanisms available for ammonia-methanol blends. Among the five results calculated using M5 are consistently more skewed than the re­
mechanisms, only M2 and M3 are dedicatedly developed for ammonia- sults of the CV model. For A95, only M3 successfully calculated the
methanol blends, while the other three mechanisms are not. The ignition delay times using the VH model, while other mechanisms failed
applicability of the five mechanisms to this study will be evaluated in the to yield valid results, because the calculated ignition delay times are
next section. longer than 1 s.

3. Results and discussion 3.1.2. Species sampling results


A80 and A95 were selected for sampling experiments as the
3.1. Experimental results and mechanism validation maximum pressures in the combustion chamber for the other two fuel
blends were much higher than the threshold of the sampling valve. In
3.1.1. Ignition delay time such cases, a too high pressure could make the sampling valve open for
Fig. 4 illustrates the experimental and the calculated results of the too long or directly damage the sampling valve. In a sampling experi­
ignition delay times at peff = 25 bar. The solid lines show the calculated ment, a longer ignition delay time will be affected by heat loss, while a
results using the CV model, while the dashed lines show the results using shorter ignition delay time will lead to increased uncertainty in the
the VH model. It can be observed that for A20 and A40, calculated re­ sampling process [34]. To obtain a suitable ignition delay time, sam­
sults of M5 are most consistent with the experimental results, and the pling experiments for A80 were conducted under the conditions of Teff =
results of M3 are also close to the experimental results. However, for A80 890 K and peff = 25 bar. The average ignition delay time was about 41
and A95, the results obtained from M3 are in better agreement with the ms, and the sampling resolution is 10.25:1 (average ignition delay 41
experimental results. The calculated results using M5 are also close to ms: sampling duration 4 ms). The sampling experiments for A95 were
conducted under the conditions of Teff = 932 K and peff = 25 bar. The
average ignition delay time was about 46 ms, and the sampling resolu­
Table 2 tion was 11.5:1 (average ignition delay 46 ms: sampling duration 4 ms).
Reaction mechanisms for ammonia-methanol oxidation simulation.
It should be noted that the mole fraction calculated from the GC
No. Mechanism Source deviates from the real value of the sampled species due to the existence
M1 Dai et al. Ref. [9] of a dead volume (the volume of the sampling probe) in the sampling
M2 Wang et al. Ref. [25] system. The unreacted gas in the dead volume dilutes the sampled
M3 Li et al. Ref. [29] mixture from the ’adiabatic core region’. A correction coefficient ξ
M4 Li et al. Ref. [36]
defined by Eq. (4) is calculated to compensate for the above-mentioned
M5 Glarborg et al. Ref. [37]

4
Q. Zhang et al. Fuel 368 (2024) 131636

Fig. 4. Comparison of experimental and calculated results of ignition delay time at peff = 25 bar.

dilution effect, where pdead, Tdead and Vdead represent the pressure, uncertainty of the measurement of the ignition delay time, respectively.
temperature and volume of the dead volume, respectively. psampling, The detailed calculation method for δτign can be found in Ref. [39,40].
Tsampling and Vsampling represent the pressure, temperature and volume of The obtained uncertainties of sampling times are shown in Fig. 5 and
the sampling chamber, respectively. The detailed calculation procedure Fig. 6.
of ξ and its uncertainty can be found in Ref. [33,38] and the results are
δτsampling τ
shown in Fig. 5 and Fig. 6. δτ = − ⋅δτ (6)
τign τign ign
The real value of species concentration in ’adiabatic core region’, χ i,
core, was calibrated by Eq. (5) using the original value χ i,samping (calcu­ Based on the calculated results of the ignition delay time, mecha­
lated by GC signal). For all species except the initial reactants ammonia nisms M3 (Li et al.) and M5 (Glarborg et al.) were selected to numeri­
and methanol involved in this work, χ i,dead is 0, and the χ i,dead of cally calculate the evolution of intermediate species concentration. The
ammonia and methanol is the same as the concentration in the initial results, as shown in Fig. 5 and Fig. 6, indicate that both the two mech­
mixture. anisms can approximately predict the trend of intermediate species
( )− 1 concentration, and M3 has better performance. However, both the two
pdead ⋅Vdead psampling ⋅Vsampling mechanisms exhibit noticeable deviations in predicting the final con­
ξ= ⋅ (4)
Tdead Tsampling sumption of CO and the final production of CO2, which may be attrib­
uted to the incomplete consumption path of CO in the two mechanisms.
χ i,core *(1 − ξ) + χ i,dead *ξ = χ i,sampling (5)
These discrepancies do not have an impact on the analysis of the con­
The sampling results were averaged since the sampling valve opened sumption of fuels and their interactive reactions during the ignition
for 4 ms. And the initial temperature fluctuation and partial pre-reaction process, which are the focus of this study. Therefore, subsequent nu­
in the compression process will affect the measurement results of the merical calculations are based on M3 due to its accuracy in predicting
ignition delay time. Therefore, it is necessary to consider the uncertainty both the ignition delay time and intermediate species.
of the sampling time. This can be determined using the method
described in Ref. [39], as shown in Eq. (6). In this equation, δτsampling,
τign, τ, and δτign represents half of the sampling valve opening duration,
the overall ignition delay time, the normalized sampling timing, and the

5
Q. Zhang et al. Fuel 368 (2024) 131636

Fig. 5. Sampling results and simulation results of the intermediate species of A80.

3.2. Ignition-enhancing effect of methanol on ammonia sensitive reaction for OH radical generation, except for R21 of H2O2
decomposition to OH. Even for A95, which has the smallest proportion
3.2.1. Reduction of ignition delay time of methanol, R119 and R123 remain among the top five reactions with
The ignition delay times of ammonia-methanol blended fuel under relatively large sensitivity coefficients. As the proportion of ammonia
four ratios are shown in Fig. 7. It is evident that with an increase in the decreases, the influence of ammonia-related reactions on OH radical
proportion of methanol, the ignition delay time is significantly reduced, generation gradually decreases. In the case of A40, only one of the first
suggesting that the addition of methanol improves the reactivity of the five reactions (R1015) is related to ammonia, with a very small sensi­
fuel. However, when the proportion of methanol surpassed 20 %, there tivity coefficient. For A20, there are no reactions directly related to
is little variation in the ignition delay times, consistent with the results ammonia among the first five reactions. This means that methanol
reported by Li et al. [29]. significantly enhances the production of OH radical, thereby boosting
It is worth noting that pure ammonia did not ignite in the tempera­ the overall reactivity of the mixture.
ture range of 810 K-970 K and the lowest ignition temperature of pure
ammonia was about 1050 K. This indicates that adding methanol is 3.3. Fuel consumption patterns
efficient in lowering the minimum ignition temperature of ammonia.
Different fuel consumption patterns of A80 and A95 were observed
3.2.2. ROP and sensitivity of OH radical from the sampling results. As shown in Fig. 5, for A80, in the early stage
Given the importance of the OH radical accumulation in te ignition of the ignition process (normalized timing < 0.8), methanol consumes
process and its significant contribution to the consumption of methanol slowly while ammonia shows almost no consumption. When the
and ammonia [29], the rate of production (ROP) of OH radical was normalized timing is close to 1, most of both ammonia and methanol
calculated using M3, as shown in Fig. 8. The calculation conditions for were rapidly consumed. The intermediate species CO and N2O begin to
the four blended fuels were Teff = 890 K and peff = 25 bar. It can be accumulate slowly at the normalized time of around 0.6, with peak
observed that an increase in methanol proportion results in an increase concentrations appearing just before the ignition timing (normalized
in the peak value of both the OH fraction and its ROP, demonstrating timing of around 0.9), and then are rapidly consumed. The majority of
enhanced reactivity of the mixture during the ignition process. final products CO2 and N2 are generated at the moment of ignition, with
The sensitivity of OH radical at the ignition timing were calculated to very little production in the early stage of the ignition process. For A95,
further analyze the leading reactions controlling its generation. The top as shown in Fig. 6, methanol begins to be consumed at a certain rate
five reactions most sensitive to the generation of OH radical are selected from the EOC to the ignition timing, without any obvious instantaneous
to be shown in Fig. 9. The sensitivity results indicate the significant role consumption process. This may be attributed to the reduction of the
of methanol reactions in OH radical generation. When the proportion of initial concentration of methanol. However, ammonia experiences slow
methanol is 20 % or more (A80, A40, A20), R123 is always the most consumption in the early stage of the ignition process and then is rapidly

6
Q. Zhang et al. Fuel 368 (2024) 131636

Fig. 6. Sampling results and simulation results of the intermediate species of A95.

Fig. 7. Experimental results of ignition delay time.

consumed when close to the ignition timing. Different from A80, the M3 was used to calculate the ROP of NH3, CH3OH and their inter­
intermediate species CO and N2O of A95 are slowly generated from the mediate species to further analyze the reasons for the different fuel
EOC, then the generation rate increases slightly after the normalized consumption rates of A80 and A95. Li et al. pointed out that the inter­
time of around 0.6. Finally, CO and N2O are consumed rapidly at the action between NH2 and CH2O has a significant inhibitory effect on the
moment of ignition. The generation mode of final products CO2 and N2 consumption of NH3 [29]. Hence, the ROP of CH3OH, CH2O, NH3 and
are similar to that of A80. NH2 were examined, as illustrated in Fig. 10 and Fig. 11. The five

7
Q. Zhang et al. Fuel 368 (2024) 131636

Fig. 8. Molar fraction and ROP of OH radical.

Fig. 9. Sensitivity to OH radical formation.

reactions listed in the figures are the ones with the highest ROP from the via R130. The most important consumption reaction of CH2O is R1357,
EOC to the ignition timing. Dash lines in the figures indicate reactions in which CH2O is further dehydrogenated to HCO, reducing NH2 to NH3,
involving both carbon and nitrogen species, while solid lines represent which indicates that NH2 promotes the consumption of CH3OH. That is
reactions involving only carbon or nitrogen species. why CH3OH begin to be consumed from the early stage of the ignition
The ROP of key reactions in Fig. 10 indicates that NH2 can generate process for A80.
from NH3 by R1015, but because of the presence of CH3OH, the inter­ For A95, the ROP of key reactions in Fig. 11 indicates that most of
mediate species CH2O continuously reduces NH2 to NH3 by R1537, NH3 converts to NH2 through R1015. Although there are still reactions
which significantly inhibits the consumption of NH3. That is why NH3 (R1017 et al.) inhibiting NH3 consumption, these reactions are inde­
does not consume in the early stage for A80 as mentioned above. R1017 pendent of carbon-containing species, and their rates are very low. The
also inhibits NH3 consumption, but this reaction is independent of major consumption reactions (R1023, etc.) of NH2 are also independent
carbon-containing species and the reaction rate is much less than that of of carbon-containing species. For CH3OH, no reactions impede its con­
R1537. Meanwhile, no reaction inhibits the consumption of CH3OH, sumption, similar to that for A80. The main consumption reactions are
which mainly converts to CH2OH via R119, and further generates CH2O R119, R1793 and R1794. R1793 and R1794 are interactive reactions

8
Q. Zhang et al. Fuel 368 (2024) 131636

Fig. 10. ROP of key reactions of some species for A80.

Fig. 11. ROP of key reactions of some species for A95.

between CH3OH and NH2, where CH3OH dehydrogenates, and NH2 is inhibits the consumption of NH3. However, the lower reaction rates of
reduced to NH3. Therefore, in the case of A95, NH2 still promotes the R1793 and R1794 compared to other reactions of NH3 limit their
consumption of CH3OH while the presence of CH3OH, in turn, still inhibitory effect on NH3 consumption. As for the consumption reactions

9
Q. Zhang et al. Fuel 368 (2024) 131636

of CH2O, R1357 exhibits the highest reaction rate, which involves NH2. respectively. The corresponding reaction fluxes are presented in Fig. 12
CH2O dehydrogenates and converts to HCO while NH2 is reduced to and Fig. 13, with colored reaction paths for clear distinction. Carbon-
NH3, further supporting the conclusion that NH2 promotes the con­ only reaction paths were depicted by brown arrows, nitrogen-only re­
sumption of CH3OH. However, R1357 is not dominant for NH2 and NH3, action paths by blue arrows, and carbon–nitrogen interaction paths by
as illustrated in Fig. 11. The inhibitory effect of CH3OH on NH3 con­ orange arrows. The species annotated alongside the arrows are the re­
sumption is not apparent. That is why both NH3 and CH3OH are slowly actants of the species at the start of the arrows. The contribution of each
consumed from the early stage of the ignition process for A95 as species is indicated by the percentage number, with different colors
mentioned above. representing different proportions of fuel. The dashed arrows indicate
In summary, the consumption of ammonia was significantly inhibi­ that the species at the start is insignificant (the reaction flux of upstream
ted by methanol and its intermediate species CH2O at a methanol pro­ species is less than 5 %) at low methanol proportions but becomes sig­
portion of 20 % (A80). Conversely, the consumption of methanol was nificant at high methanol proportions. It is worth noting that the reac­
promoted by ammonia and its intermediate species NH2. At a lower tion paths of nitrogen-containing species that does not participate in the
methanol ratio (A95), ammonia still promoted the consumption of interaction are not included in Fig. 12 and Fig. 13 since this study fo­
methanol. However, although methanol and CH2O still inhibited the cuses on the consumption of fuel and the interaction between carbon-
consumption of ammonia, this inhibitory effect was not evident due to containing and nitrogen-containing species.
the low concentration of methanol. Consistent with the conclusion in Section 3.3, Fig. 12 reveals that in
the early stage of the reaction (when methanol consumption is 50 %),
the consumption of ammonia is inhibited. This inhibition is manifested
3.4. Interactive reaction paths between ammonia and methanol by methanol and its intermediate species CH2O reducing NH2 to
ammonia. With the increase of methanol ratio, the reduction path of
Analyses of reaction flux were conducted for the four ammonia- NH2 to NH3 by CH2O is gradually enhanced. With an increase in the
methanol blends based on the ROP calculated using M3. This aimed to proportion of ammonia, NH2 will promote the consumption of meth­
further compare the reaction paths of ammonia, methanol, and their anol, and the reactions of methanol and CH2O with NH2 gradually
intermediate species. For consistency with experimental conditions, the dominate. The contribution of the reaction of methanol with NH2 in­
calculation conditions for A95 were Teff = 932 K and peff = 25 bar while creases from < 5 % to 41.3 % and the reaction of CH2O with NH2 in­
the conditions of the other three kinds of blends were Teff = 890 K and creases from 14.5 % to 99.1 %. Conversely, the reaction between NH2
peff = 25 bar. It was found that there was minimal impact on the reaction and methanol weakens as the proportion of ammonia decreases, but
flux when the temperature changed from 890 K to 932 K for A95 fuel. methanol can still be consumed by the reaction with OH radical. Addi­
The timing of 50 % methanol consumption and ignition timing were tionally, an interaction is observed between CO and N2O with both
selected to represent the early and late stage of the ignition process,

Fig. 12. Integrated reaction flux at the timing of 50% methanol consumption.

10
Q. Zhang et al. Fuel 368 (2024) 131636

Fig. 13. Integrated reaction flux at the ignition timing.

species contributing to each other’s consumption. This interaction is was little difference in ignition delay times when the proportion
enhanced with the increase of the proportion of ammonia. of methanol exceeded 20 %. The peak value of OH radical
The interaction between ammonia and methanol described previ­ increased with methanol proportion, indicating the significant
ously persists until the ignition timing, as depicted in Fig. 13. The dif­ effect of methanol on the overall reactivity of the mixture.
ference is that, at the ignition timing, the significance of CH2O-related (2) Methanol and ammonia showed different consumption patterns
reactions decreases (<5%) for NH2 when the proportion of methanol is at different mixing ratios. For A80, methanol started to consume
low (A95). In contrast, the reaction between NH2 and HO2 intensifies, slowly at the early stage of the ignition process, while ammonia
which also inhibits the consumption of ammonia. For A80 and A95, showed no consumption. This suggests significant inhibition of
HNCO and NCO are generated. The generation of HNCO is a result of the ammonia by methanol and its intermediate species CH2O, as
reaction between CO and NH2. The subsequent evolution path of HNCO evidenced by the ROP analysis. As the ignition timing
and NCO leads to the final production of CO2 and N2. Therefore, the approached, both methanol and ammonia were consumed
interaction between CO and NH2 promoted fuel consumption, and this rapidly. However, for A95, methanol consumed at a relatively
interaction was enhanced with the increase of ammonia proportion. steady rate from the EOC all the way to the ignition timing,
However, it is crucial to note the toxicity of HNCO and NCO. If the re­ without any obvious instantaneous consumption process. In
actions are incomplete, they may cause more severe emission problems. contrast, ammonia experienced slow consumption during the
early stage of the ignition process, which is due to the weakened
4. Conclusions inhibition effect of methanol and CH2O.
(3) The consumption of ammonia in both early and late stages of the
This study examined the ignition characteristics of ammonia- ignition process was inhibited by methanol and its intermediate
methanol blended fuels with four mixing ratios under stoichiometric species CH2O, and the reduction path of NH2 to NH3 by CH2O was
conditions at peff = 15/20/25 bar and TEOC = 810–970 K. The key gradually enhanced with increasing methanol proportion. As the
findings from experiments and simulations can be summarized as proportion of ammonia increased, NH2 would promote the con­
follows: sumption of methanol. Additionally, it was observed that CO
interacted with NH2 and N2O, which promoted the consumption
(1) The addition of methanol lowered the ignition temperature and of each other. This interaction was enhanced with increasing
shortened the ignition delay time of ammonia. However, there ammonia proportion.

11
Q. Zhang et al. Fuel 368 (2024) 131636

CRediT authorship contribution statement [17] Liao W, Chu Z, Wang Y, Li S, Yang B. An experimental and modeling study on auto-
ignition of ammonia in an RCM with N2O and H2 addition. Proc Combust Inst
2023;39(4):4377–85.
Qihang Zhang: Writing – original draft, Software, Methodology, [18] Xu H, Wang J, Zhang C, Dai L, He Z, Wang Q. Numerical study on laminar burning
Investigation, Formal analysis, Data curation. Ridong Zhang: Writing – velocity of ammonia flame with methanol addition. Int J Hydr Energy 2022;47
review & editing, Investigation. Yunliang Qi: Writing – review & (65):28152–64.
[19] Gross CW, Kong S-C. Performance characteristics of a compression-ignition engine
editing, Validation, Software, Methodology, Formal analysis. Zhi Wang: using direct-injection ammonia–DME mixtures. Fuel 2013;103:1069–79.
Conceptualization, Project administration, Supervision. [20] Zhang B, Chen Y, Jiang Y, Lu W, Liu W. Effect of compression ratio and Miller cycle
on performance of methanol engine under medium and low loads. Fuel 2023;351.
[21] Feng H, Chen X, Sun L, Ma R, Zhang X, Zhu L, et al. The effect of methanol/diesel
Declaration of competing interest fuel blends with co-solvent on diesel engine combustion based on experiment and
exergy analysis. Energy 2023;282.
The authors declare that they have no known competing financial [22] Örs İ, Yelbey S, Gülcan HE, Sayın Kul B, Ciniviz M. Evaluation of detailed
combustion, energy and exergy analysis on ethanol-gasoline and methanol-
interests or personal relationships that could have appeared to influence gasoline blends of a spark ignition engine. Fuel 2023;354.
the work reported in this paper. [23] Hasan AO, Osman AI, AaH A-M, Al-Rawashdeh H, Abu-jrai A, Ahmad R, et al. An
experimental study of engine characteristics and tailpipe emissions from modern
DI diesel engine fuelled with methanol/diesel blends. Fuel Process Technol 2021;
Data availability 220.
[24] Sun Y, Qian X, Yuan M, Zhang Q, Li Z. Investigation on the explosion limits and
Data will be made available on request. flame propagation characteristics of premixed methanol-gasoline blends. Case Stud
Therm Eng 2021;26.
[25] Wang Z, Han X, He Y, Zhu R, Zhu Y, Zhou Z, et al. Experimental and kinetic study
Acknowledgment on the laminar burning velocities of NH3 mixing with CH3OH and C2H5OH in
premixed flames. Combust Flame 2021;229.
This study was supported by the National Natural Science Founda­ [26] Li X, Ma Z, Jin Y, Wang X, Xi Z, Hu S, et al. Effect of methanol blending on the high-
temperature auto-ignition of ammonia: an experimental and modeling study. Fuel
tion of China (Grant Nos.: T2341002, T2241003, and 52076118) and 2023;339.
Tsinghua-Toyota Joint Research Fund (Grant No.: 20213930001). [27] Lu M, Dong D, Wei F, Long W, Wang Y, Cong L, et al. Chemical mechanism of
ammonia-methanol combustion and chemical reaction kinetics analysis for
different methanol blends. Fuel 2023;341.
References [28] Zhuang Y, Wu R, Wang X, Zhai R, Gao C. An experimental and modeling study on
the oxidation of ammonia-methanol mixtures in a jet stirred reactor. Fuel 2024;
[1] Xing H, Stuart C, Spence S, Chen H. Alternative fuel options for low carbon 356.
maritime transportation: pathways to 2050. J Clean Prod 2021;297. [29] Li M, He X, Hashemi H, Glarborg P, Lowe VM, Marshall P, et al. An experimental
[2] Kalghatgi G, Levinsky H, Colket M. Future transportation fuels. Prog Energy and modeling study on auto-ignition kinetics of ammonia/methanol mixtures at
Combust Sci 2018;69:103–5. intermediate temperature and high pressure. Combust Flame 2022;242.
[3] Qi Y, Liu W, Liu S, Wang W, Peng Y, Wang Z. A review on ammonia-hydrogen [30] Zhang R, Liu W, Zhang Q, Qi Y, Wang Z. Auto-ignition and knocking combustion
fueled internal combustion engines. eTransportation 2023;18. characteristics of iso-octane-ammonia fuel blends in a rapid compression machine.
[4] Giddey S, Badwal SPS, Munnings C, Dolan M. Ammonia as a renewable energy Fuel 2023;352.
transportation media. ACS Sustainable Chem Eng 2017;5(11):10231–9. [31] Di H, He X, Zhang P, Wang Z, Wooldridge MS, Law CK, et al. Effects of buffer gas
[5] Valera-Medina A, Xiao H, Owen-Jones M, David WIF, Bowen PJ. Ammonia for composition on low temperature ignition of iso-octane and n-heptane. Combust
power. Prog Energy Combust Sci 2018;69:63–102. Flame 2014;161(10):2531–8.
[6] Mørch CS, Bjerre A, Gøttrup MP, Sorenson SC, Schramm J. Ammonia/hydrogen [32] Sung C-J, Curran HJ. Using rapid compression machines for chemical kinetics
mixtures in an SI-engine: Engine performance and analysis of a proposed fuel studies. Prog Energy Combust Sci 2014;44:1–18.
system. Fuel 2011;90(2):854–64. [33] Zhang P, Ji W, He T, He X, Wang Z, Yang B, et al. First-stage ignition delay in the
[7] Li J, Zhang R, Pan J, Wei H, Shu G, Chen L. Ammonia and hydrogen blending negative temperature coefficient behavior: Experiment and simulation. Combust
effects on combustion stabilities in optical SI engines. Energy Convers Manage Flame 2016;167:14–23.
2023;280. [34] Liu W, Qi Y, Zhang R, Zhang Q, Wang Z. Hydrogen production from ammonia-rich
[8] Dai L, Gersen S, Glarborg P, Mokhov A, Levinsky H. Autoignition studies of NH3/ combustion for fuel reforming under high temperature and high pressure
CH4 mixtures at high pressure. Combust Flame 2020;218:19–26. conditions. Fuel 2022;327.
[9] Dai L, Gersen S, Glarborg P, Levinsky H, Mokhov A. Experimental and numerical [35] Liu W, Qi Y, Zhang R, Wang Z. Flame propagation and auto-ignition behavior of
analysis of the autoignition behavior of NH3 and NH3/H2 mixtures at high iso-octane across the negative temperature coefficient (NTC) region on a rapid
pressure. Combust Flame 2020;215:134–44. compression machine. Combust Flame 2022;235.
[10] Okafor EC, Naito Y, Colson S, Ichikawa A, Kudo T, Hayakawa A, et al. Experimental [36] Li M, Zhu D, He X, Moshammer K, Fernandes R, Shu B. Experimental and kinetic
and numerical study of the laminar burning velocity of CH4–NH3–air premixed modeling study on auto-ignition properties of ammonia/ethanol blends at
flames. Combust Flame 2018;187:185–98. intermediate temperatures and high pressures. Proc Combust Inst 2023;39(1):
[11] Dai L, Hashemi H, Glarborg P, Gersen S, Marshall P, Mokhov A, et al. Ignition delay 511–9.
times of NH3 /DME blends at high pressure and low DME fraction: RCM [37] Glarborg P, Miller JA, Ruscic B, Klippenstein SJ. Modeling nitrogen chemistry in
experiments and simulations. Combust Flame 2021;227:120–34. combustion. Prog Energy Combust Sci 2018;67:31–68.
[12] Zhang M, An Z, Wang L, Wei X, Jianayihan B, Wang J, et al. The regulation effect of [38] Liu C, Wang Z, Song H, Qi Y, Li Y, Li F, et al. Experimental and numerical
methane and hydrogen on the emission characteristics of ammonia/air combustion investigation on H2/CO formation and their effects on combustion characteristics
in a model combustor. Int J Hydr Energy 2021;46(40):21013–25. in a natural gas SI engine. Energy 2018;143:597–605.
[13] Meng X, Zhang M, Zhao C, Tian H, Tian J, Long W, et al. Study of combustion and [39] Zhang P. Rapid compression machine coupled with sampling system: an
NO chemical reaction mechanism in ammonia blended with DME. Fuel 2022;319. experimental determination of reaction rate in combustion. Tsinghua University
[14] Schaefer DX, Jianzhong; Vogt, Mathias; Kamps, Alvaro Perez-Salado; Maurer, 2018.
Gerd. Experimental investigation of the solubility of ammonia in methanol. J Chem [40] Ji W, Zhang P, He T, Wang Z, Tao L, He X, et al. Intermediate species measurement
Eng Data 2007. during iso-butanol auto-ignition. Combust Flame 2015;162(10):3541–53.
[15] IRENA. Innovation Outlook: Renewable Methanol. In: Institute M, ed.; 2023.
[16] Santasalo-Aarnio A, Nyari J, Wojcieszyk M, Kaario O, Kroyan Y, Magdeldin M,
et al. Application of synthetic renewable methanol to power the future propulsion.
SAE Int 2020.

12

You might also like